Sunteți pe pagina 1din 7

Electrochimica Acta 56 (2011) 20242030

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Preparation and characterization of coreshell structured catalysts using Ptx Pdy as active shell and nano-sized Ru as core for potential direct formic acid fuel cell application
Haili Gao, Shijun Liao, Jianhuang Zeng , Yichun Xie, Dai Dang
School of Chemistry and Chemical Engineering, South China University of Technology, 381 Wushan Road, Tianhe District, Guangzhou 510641, Guangdong, PR China

a r t i c l e

i n f o

a b s t r a c t
Carbon-supported coreshell structured Ru@Ptx Pdy /C catalysts with Ptx Pdy as shell and nano-sized Ru as core are prepared by a successive reduction procedure. The catalysts are extensively characterized by transmission electron microscopy (TEM), X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). The formic acid oxidation activity of Ru@Ptx Pdy /C varies with the varying Pt:Pd atomic ratio. The peak oxidation potential on Ru@Pt1 Pd2 /C shifts negatively for about 200 mV compared with that of Pd/C. The higher electro-catalytic activity toward formic acid oxidation on coreshell structured Ru@Ptx Pdy /C catalyst than that on Ptx Pdy /C suggests the high utilization of noble metals. In addition to the enhanced noble metal utilization, Ru@Ptx Pdy /C catalyst also shows improved stability as evidenced by chronoamperometric evaluations. 2010 Elsevier Ltd. All rights reserved.

Article history: Received 19 September 2010 Received in revised form 23 November 2010 Accepted 26 November 2010 Available online 3 December 2010 Keywords: Ru@Ptx Pdy /C Catalysts Coreshell structure Formic acid oxidation

1. Introduction In view of the advantages over direct methanol fuel cells including higher theoretical open circuit potential (1.45 V) [1], lower fuel crossover [2], less positive oxidation potential [3] and non-toxicity, direct formic acid fuel cells (DFAFCs) have been studied extensively and were considered as candidates for portable electronic devices such as cellular phones, personal digital assistants (PDAs) and laptop computers. It has been believed that platinum shows the highest catalytic activity for electro-oxidation of formic acid relative to palladium, rhodium and gold electrodes [4]. The widely accepted reaction mechanism for formic acid oxidation on Pt is dual path mechanism [5], which involves a reactive intermediate formed through a main dehydrogenation path and adsorbed CO (poisoning species) formed in a parallel dehydration path. The adsorption of the intermediate CO hampers the catalyst activity signicantly [6]. In addition, high cost and resource scarcity of Pt poses another limitation to the commercialization of DFAFCs. In order to achieve the aims of enhanced activity and reduced cost, extensive efforts have been made to enhance the oxidation rate of formic acid on Pt catalysts by alloying Pt with Ag [7], Fe [4], Pb [8], etc. Recently Masel and co-workers [9,10] discovered that Pd and Pd/C catalysts are less suffered by CO poisoning because the electro-oxidation of formic

Corresponding author. Tel.: +86 20 8711 3586; fax: +86 20 8711 3586/2977. E-mail address: cejhzeng@scut.edu.cn (J. Zeng). 0013-4686/$ see front matter 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.electacta.2010.11.090

acid follows the direct pathway on them. Unfortunately Pd alone catalysts are not the nal solutions for formic acid oxidation. Pd is prone to oxidation compared with Pt [11], therefore the stability of Pd catalysts are questionable [12]. It was reported that Pd shows much higher start-up formic acid oxidation activity than that of Pt or Pt/Pd catalyst, but the steady-state activities are in the order of Pt/Pd > Pd > Pt [1,13]. Exploration of foreign elements addition to palladium is therefore extensively attempted to netune the Pd activity through the means of PtPd single crystal [14], Pd decorated Pt [15] and palladium-coated Pt [16]. Coreshell conguration construction is certainly an effective way to increase the electro-catalytic activity of catalysts and to reduce the catalyst cost. Lee and co-workers studied coreshell structured Au@Pd catalyst and it was found that the catalytic activity and stability of Pd for formic acid oxidation could be substantially improved due to the interaction between Pd shell and Au core [17]. Jiang et al. prepared platinum sub-monolayer decorated gold nanorods and applied for formic acid oxidation and it was found this catalyst is superior to conventional Pt/C catalyst. They attributed the enhanced performance to the suppression of poisoning intermediate CO formation on the tailored catalyst [18]. Very recently, research reported from our lab identied Pt@Pd/C as ideal electrocatalyst for anode oxidation of formic acid owing to the high dispersion of nanoparticles and synergistic effect between Pt and Pd at the interface [19]. Herein, we report the preparation and characterizations of coreshell structured Ru@Ptx Pdy /C nanoparticles for formic acid electro-oxidation. Ruthenium was selected as the core element

H. Gao et al. / Electrochimica Acta 56 (2011) 20242030

2025

Intensity / a.u.

because very small and highly dispersed Ru nanoparticles can be easily prepared, hopefully yielding highly stable and high noble metal utilization catalyst. Another gained advantage for the usage of Ru core is the expected catalytic activity enhancement between Ru and Pt due to the modied reaction kinetics [20]. Electrochemical evaluation by cyclic voltammetry, indeed, conrmed the benecial effect of Ru core and Ru@Ptx Pdy /C registered high catalytic activity for formic acid oxidation. 2. Experimental 2.1. Catalyst preparation

#: Ru : PtPd alloy

A
#

Ru@Pt1Pd2/C Ru@Pt1Pd1/C Ru@Pt2Pd1/C Pt2Pd1/C Ru/C

Carbon supported Ru@Ptx Pdy /C catalysts with Ptx Pdy alloys on the surface of Ru cores were prepared by a successive reduction route. Preparation of Ru/C: XC-72 carbon powders were pretreated by 10% HNO3 and 30% H2 O2 at 80 C. 207 mg RuCl3 3H2 O was dissolved in diluted HCl solution, followed by the addition of 300 mg as-prepared carbon powders. The mixture was then allowed for evaporation in a water bath at 70 C. Reduction of ruthenium chloride on carbon (Ru/C) was completed by heating the dried power in a ceramic boat at 200 C under owing H2 for 2 h. Preparation of Ru@Ptx Pdy /C catalysts: appropriate amounts of palladium chloride (PdCl2 ) solution, platinum chloride (H2 PtCl6 6H2 O) solution and sodium citrate were dissolved in 15 ml ethylene glycol (EG) and then stirred for 1 h to entirely dissolve sodium citrate. Afterwards, as-prepared Ru/C was added to the mixture, followed by pH adjustment to >10 by drop-wise addition of 5 wt.% KOH/EG solution under vigorous stirring. The mixture was then transferred into a Teon-lined autoclave and conditioned at 130 C for 6 h, followed by ltering, washing and vacuum drying at 70 C. For comparison, a 20 wt.% Pt/C, 20 wt.% Pd/C and 20 wt.% Pt2 Pd1 /C (atom ratio of PdPt was 1:2) catalysts were prepared by a similar way as that of Ru@Ptx Pdy /C, except that C was added instead of Ru/C. The atom ratio of Pd to Pt was 2:1, 1:1 and 1:2 in each sample. Ru@Pt2 Pd1 /C, Ru@Pt1 Pd1 /C and Ru@Pt1 Pd2 /C denote the atomic ratios of Ptx :Pdy are 2:1, 1:1 and 1:2, respectively. 2.2. Materials characterization X-ray powder diffraction (XRD) patterns of the catalysts were recorded by a Shimadzu XD-3A (Japan), using ltered Cu K radiation at 35 kV and 30 mA. The 2 angles were scanned from 20 to 80 at 4 min1 and a step size of 0.01 . The particle size and surface morphologies of the catalysts were analyzed via transmission electron microscopy (JEOL JEM-2010HR, Japan) operated at 200 kV. X-ray photoelectron spectroscopy (XPS) measurement was carried out on a Perkin Elmer PHI1600 system (PerkinElmer, USA) using a single Mg-K radiation source operating at 300 W and 15 kV. Spectral correction was based on the C1s level of graphitic carbon at 284.5 eV. 2.3. Electrochemical evaluations An electrochemical workstation (Ivium, Netherlands) and a three-electrode electrochemical cell were used for the electrochemical measurements. The working electrode was a thin layer of Naon-impregnated catalyst ink cast on a 5 mm diameter glassy carbon electrode: the ink was made by dispersing 5 mg catalyst ultrasonically in 1 ml Naon/ethanol (0.25 wt.% Naon) for 30 min; 6 l of the ink was then pipetted and spread on the glassy carbon surface followed by dryness in the air. The counter and reference electrodes were a platinum wire and an Ag/AgCl (saturated KCl) electrode respectively. All potentials presented are quoted with

20

30

40

50

60

70

80

Two theta / degree

B
Pt2Pd1/C

Intensity / a.u.

Ru@Pt2Pd1/C Ru/C

60

63

66

69

72

75

Two theta / degree


Fig. 1. X-ray diffraction patterns of Ru@Ptx Pdy /C, Pt2 Pd1 /C and Ru/C catalysts (A) and the zoomed view of peak prole for Ru@Pt2 Pd1 /C between 60 and 75 (B).

respect to the normal hydrogen electrode (NHE). Cyclic voltammetry curves were obtained in 0.50 M H2 SO4 for electrochemically active surface area measurements and the catalytic activity evaluation was carried out in 0.50 M H2 SO4 plus 2.0 M HCOOH solution at a scan rate of 30 mV s1 . The chronoamperometric experiments were measured at 0.3 V for 1000 s in the same electrolyte. The solutions were purged with high purity N2 for 20 min prior to each measurement. 2.4. Single cell test Membrane electrode assembly, each with an active area of 4 cm2 , was fabricated using the catalyst-sprayed membrane method [21]. Catalyst inks were prepared by dispersing catalyst powder into a mixture of isopropanol (Sinopharm Chemical Reagent Co.) and 5 wt.% Naon ionomer solution (DuPont, USA, 5 wt.% in alcohol). The dispersion mixture was ultrasonicated for 30 min before use. The inks were then sprayed onto each side of Naon 212 membrane (Dupont) using an air brush. The MEA was vacumm-dried at 70 C overnight before usage. Pt/C (Johnson Matthey, 40 wt.%) was used as cathode catalyst with a loading of 1 mg Pt cm2 . Pt/C (Johnson Matthey, 40 wt.%), Pt2 Pd1 /C or Ru@Pt2 Pd1 /C catalysts were used as anode catalysts respectively with a loading of 1 mg Pt cm2 or 1 mg PtPd cm2 . Carbon papers (TORAY) used as diffusion layers without any pretreatment were sandwiched between the catalyst layers and the current collectors. The cell IV curves were obtained using a home-made testing device. External resistor (11000 ) was used for adjustment of the resistance. The corresponding current and voltage were recoded

2026

H. Gao et al. / Electrochimica Acta 56 (2011) 20242030

Fig. 2. TEM images of the as-prepared Ru/C (A and C), Ru@Pt2 Pd1 /C (B and D) and the corresponding particle size distribution of Ru/C (E) and Ru@Pt2 Pd1 /C (F).

5 s after the resistance was set. This allows for sufcient time for the voltage to reach stable status. The constant current test was acquired through an electrochemical workstation (eDAQ data acquisition systems) using galvanostatic testing mode. Formic acid (88% Analytical grade, Enox ) was diluted with deionization water to give a concentration of 5.0 M. 3. Results and discussion The XRD patterns of Ru@Ptx Pdy /C, Pt2 Pd1 /C and Ru/C catalysts are shown in Fig. 1A. The peak centering at 24.5 for all the cata-

lysts can be ascribed to the graphitic nature of the carbon support. There are three observable peaks at 2 angles of ca. 40 , 47 and 68 for Ru@Pt2 Pd1 /C, Ru@Pt1 Pd1 /C and Ru@Pt1 Pd2 /C, respectively. The only subtle difference for the XRD pattern of the catalysts is the peak position, which is resulted from the crystallite deformation of Ptx Pdy alloy with variable Pt:Pd atomic ratio (the lattice mismatch of Pt and Pd is very small, at 0.7%). Compared with Ru/C, the characteristic peaks of ruthenium in Ru@Ptx Pdy /C are almost disappeared except for a weak Ru (1 0 1) reection peak located at ca. 44.0 . This is probably due to the covering of Ptx Pdy alloy on Ru surface, leading to dwarfed peak intensity. The reduction of Ptx Pdy

H. Gao et al. / Electrochimica Acta 56 (2011) 20242030

2027

A
Intensity / cps

Pt 4f5/2

Pt 4f7/2

B
Intensity / cps

Pt 4f5/2

Pt 4f7/2

Pt(0) Pt(0) Pt(2) Pt(2)

Pt(0) Pt(2) Pt(2) Ru@Pt1Pd1/C Pt(0)

Ru@Pt1Pd2/C

76

74

72

70

76

74

72

70

Binding energy / ev

Binding energy / ev
Pt 4f7/2

C
Intensity / cps

Pt 4f5/2

Pt 4f5/2

Pt 4f7/2

Pt(0) Pt(0) Pt(2) Pt(2) Ru@Pt2Pd1/C

Intensity / cps

Pt(0) Pt(2) Pt(2) Pt2Pd1/C

Pt(0)

76

74

72

70

76

74

72

70

Binding energy / ev

Binding energy / ev
Pd 3d5/2

E
Pd 3d3/2

F
Intensity / cps

Ru 3p5/2 Ru@Pt2Pd1/C

Ru 3p3/2

Intensity / cps

Ru@Pt1Pd2/C Ru@Pt2Pd1/C Ru@Pt1Pd1/C Pt2Pd1/C

Ru@Pt1Pd1/C

Ru@Pt1Pd2/C

342

340

338

336

334

500

490

480

470

460

450

Binding energy / ev

Binding energy / ev

Fig. 3. XPS spectra of Pt 4f (AD), Pd 3d (E) and Ru 3p (F) of the catalysts.

could possibly occur on the pre-formed Ru surface: in the preparation of Ru@Ptx Pdy /C, ruthenium precursor was rstly impregnated on carbon support and then reduced in H2 and the active sites of carbon support are mainly occupied by Ru particles. As a result, in the second preparation stage, the reduction of Pt and Pd is most likely to occur on the surface of Ru particles since freshly formed Ru nanoparticles have a large number of energetically unsaturated atoms. It should be noted that formation of isolated Ptx Pdy particles is certainly evitable but less likely because the creation of new nuclei is not energetically favorable in the solution phase [22]. Fig. 1B shows the zoomed view of the (2 2 0) diffraction. The average crystal size of Pt2 Pd1 /C, Ru/C and Ru@Pt2 Pd1 /C catalysts are calculated to be ca. 2.9, 2.5 and 3.0 nm, respectively. The particle size of Ru@Pt2 Pd1 /C is slightly bigger than that of Ru/C, which is possibly resulted from the decoration of Ptx Pdy alloy on the surface of the pre-formed Ru. If this were the fact, the shell thickness of Ptx Pdy could be calculated to be 0.25 nm. In addition, the (2 2 0)

diffraction of Ru@Ptx Pdy /C is located between the two theta angles of pure Ru/C and the alloyed Ptx Pdy , a phenomenon observed by other groups on coreshell structured nanoparticles [23,24]. Fig. 2 shows TEM images of Ru/C (A and C) and Ru@Pt2 Pd1 /C (B and D). In Fig. 2A and C, it can be observed that Ru particles are highly dispersed on the carbon support with narrow size distribution. The nanoparticle popularity, dened here as the number of metal nanoparticle per unit area of carbon support, seemed to be the same for Ru@Pt2 Pd1 /C except for the grown-up of the particle size due to the decoration of Ptx Pdy alloys on preformed Ru nanoparticles. By counting more than 100 particles in the image of Fig. 2C and D, the average particle size of the Ru/C and Ru@Pt2 Pd1 /C catalysts is ca. 2.5 and 3.0 nm, respectively, which is fairly agreed with the XRD observation. It is universally agreed that direct observation of coreshell structure through TEM imaging is of great difculty, otherwise, the grown-up of particle size and the unchanged nanoparticle popularity provided indirect evidence for

2028

H. Gao et al. / Electrochimica Acta 56 (2011) 20242030

3 0 -3 -6 -9 -12 0.0 0.2 0.4 0.6 0.8 1.0 1.2

Ru@Pt1Pd1/C Ru@Pt1Pd2/C Pt/C Ru@Pt2Pd1/C Pt2Pd1/C

Potential / V vs NHE
Fig. 4. Cyclic voltammograms of Ru@Ptx Pdy /C, Pt/C and Pt2 Pd1 /C catalysts in N2 purged 0.50 M H2 SO4 electrolyte at room temperature and at a sweep rate of 30 mV s1 .

the selectively deposition of Ptx Pdy on Ru surface. In fact, it is reasonable to suggest that Pt and Pd would be reduced on the small Ru particles based on the following: (i) the interaction between the metals are stronger than the interaction between metal and carbon [25]; (ii) aging effect can make small Ptx Pdy particles migrate to the surface of Ru particles [23]. The surface electronic structure of Pt2 Pd1 /C and Ru@Ptx Pdy /C catalysts was evaluated by X-ray photoelectron spectroscopy (Fig. 3). The Pt 4f spectra of the catalysts are shown in Fig. 3AD. The Pt 4f signal in all the samples can be deconvoluted into two doublets. For Pt2 Pd1 /C catalyst (Fig. 3D), the peaks with binding energies of 71.04 eV and 74.32 eV are ascribed to metallic Pt, whereas the peaks corresponding to 71.97 and 75.19 eV can be assigned to Pt (II) species in the form of Pt(OH)2 or PtO [26]. The BEs of Pt in Pt2 Pd1 /C are all negatively shifted compared with those of single Pt/C catalyst in our previous report (71.30 eV and 74.50 eV for Pt(0), 72.30 and 76.50 eV for Pt(II)) [27]. The negative shift of Pt (4f) peaks is possibly due to the addition of Pd and shows that electron transfer occurs from Pt to Pd (electronegativity: Pt = 2.28 and Pd = 2.20). The metallic Pt 4f7/2 lines for Ru@Pt1 Pd1 /C, Ru@Pt2 Pd1 /C and Ru@Pt1 Pd2 /C occurred at 71.26, 71.34, and 71.14 eV, respectively, whereas the metallic Pt 4f5/2 lines were at 74.55, 74.65, and 74.49 eV, respectively. The Pt binding energies of Ru@Ptx Pdy /C catalysts were higher than those of Pt2 Pd1 /C, indicating an interaction between Ru and Ptx Pdy . The percentage of Pt in the zero valent state for Ru@Pt1 Pd2 /C, Ru@Pt1 Pd1 /C and Ru@Pt2 Pd1 /C was higher than that in Pt2 Pd1 /C (71.2%), which was 74.6%, 75.8% and 80.4%, respectively, increasing with the increase of Pt content. The Pd 3d levels of Pt2 Pd1 /C and Ru@Ptx Pdy /C catalysts in Fig. 3E showed metallic palladium peaks only and no peaks were observed for palladium oxide [28]. The metallic Pd 3d5/2 line for Ru@Pt1 Pd1 /C, Ru@Pt2 Pd1 /C and Ru@Pt1 Pd2 /C occurred at 335.68, 335.80, and 335.62 eV, respectively, which was higher than that of Pt2 Pd1 /C catalyst (335.50 eV). The positive shift of Pd 3d and Pt 4f in the shell was well understood by other groups [19,29]. The Ru 3p spectra in Fig. 3F were laden with much noise and poorly resolved, making it difcult to quantify the BE states of Ru. However, the Ru in the Ru@Ptx Pdy /C catalysts can still be clearly identied. Cyclic voltammetry (CV) is generally regarded as a surface sensitive technique that detects electrochemical properties of surface atoms rather than bulk atoms. Fig. 4 shows the cyclic voltammograms of Pt2 Pd1 /C, Pt/C and Ru@Ptx Pdy /C catalysts in 0.50 M H2 SO4 .

Typical hydrogen adsorption and desorption region can be clearly detected on Pt/C. The hydrogen desorption feature of Ru@Pt2 Pd1 /C is similar to that of the alloyed Pt2 Pd1 /C catalyst. Both Ru@Pt1 Pd2 /C and Ru@Pt1 Pd1 /C catalysts exhibited high hydrogen peak below 0.3 V, which was probably caused by dissolution of the adsorbed hydrogen into the bulk of Pd. Higher peak for Ru@Pt1 Pd2 /C relative to Ru@Pt1 Pd1 /C conrmed the higher Pd content in the former catalyst. The electrochemical surface areas (ESAs) of different catalysts were calculated according to methods reported in the literature [30]. The specic activity is calculated to be 80.68 m2 g1 metals for Pt2 Pd1 /C and 125.05 m2 g1 metals for Ru@Pt2 Pd1 /C. The specic activity of Ru@Pt2 Pd1 /C can be converted into 250.10 m2 g1 PtPd, which is 2.1 times higher than that of Pt2 Pd1 /C. This is a suggestion of higher utilization of noble Pt and Pd on Ru@Pt2 Pd1 /C catalyst relative to Pt2 Pd1 /C. Fig. 5A shows the polarization curves of Ru@Ptx Pdy /C, Pt2 Pd1 /C, Pd/C and Pt/C electrodes for formic acid oxidation in 0.50 M H2 SO4 + 2.0 M HCOOH solution. Two oxidation peaks were observed in the forward scan for Pt2 Pd1 /C and Pt/C catalysts. The rst peak (I1 ) at ca. 0.65 V is attributed to the direct oxidation of formic acid to CO2 , whereas the second peak (I2 ) at ca. 0.95 V refers to the oxidation of the poisoning intermediate CO (generated from the dissociative adsorption step) [31]. The polarization curve pattern of Ru@Ptx Pdy /C differed from that of Pt/C or Pd/C. There are two oxidation peaks occurred at ca. 0.2 V and ca. 0.95 V for Ru@Ptx Pdy /C. Compared with Pd/C, the position of the rst peak was negatively shifted for 200 mV, an indication of enhanced activity for formic acid. The higher current density and the more negative position of the rst oxidation peak for Ru@Pt1 Pd2 /C than that of Pd/C undoubtedly showed the enhanced formic acid oxidation activity on Ru@Pt1 Pd2 /C. In addition, the rst peak intensity at ca. 0.2 V for Ru@Ptx Pdy /C increased with the increased Pd content and the second peak intensity decreased with the decreased Pd content. The peak current density at ca. 0.95 V is 16.67, 19.93, 46.09, 38.60 and 71.43 mA cm2 for Pt/C, Pt2 Pd1 /C, Ru@Pt1 Pd1 /C, Ru@Pt1 Pd2 /C and Ru@Pt2 Pd1 /C, respectively, increased in the order: Pt/C < Pt2 Pd1 /C < Ru@Pt1 Pd1 /C < Ru@Pt1 Pd2 /C < Ru@Pt2 Pd1 /C. The enhanced catalytic activity for Ru@Ptx Pdy /C is due to high Ptx Pdy utilization on highly dispersed Ru core surface. Polarization curves of Ru@Ptx Pdy /C catalysts collected from 1.2 V to 0.0 V are shown in Fig. 5B. It can be observed that there are no oxidation peaks for Ru@Ptx Pdy /C and Pt2 Pd1 /C catalysts. These catalysts are featured with a sharp current steep at 0.80.9 V. By contrast, Pt/C catalyst showed an oxidation peak at 0.7 V. It could be observed that the oxidation current intensity of Ru@Ptx Pdy /C was much stronger than that of Pt/C and Pt2 Pd1 /C, meaning the easier oxidation of formic acid and higher noble metal utilization efciency on Ru@Ptx Pdy /C. Cyclic voltammograms of the catalysts are shown in Fig. 5C. Similar to the observations in Fig. 5A and B, Ru@Ptx Pdy /C showed higher activity toward formic acid oxidation than Pt2 Pd1 /C and Pt/C. Fig. 6 shows the chronoamperometric curves of the catalysts in 0.50 M H2 SO4 + 2.0 M HCOOH at 0.3 V. The potential of the working electrode with Ru@Ptx Pdy /C and Pd/C catalysts was xed at 0.3 V and the changes in the oxidation current with time were recorded. For Pd/C catalyst, the current decayed continuously during the evaluation period, supposedly because of catalyst poisoning by the chemisorbed carbonaceous species [32]. The initial current (t < 100 s) for the catalyst was decreased according to the order: Ru@Pt2 Pd1 /C > Ru@Pt1 Pd1 /C > Ru@Pt1 Pd2 /C. The quasi-stable current density at the Ru@Pt2 Pd1 /C, Ru@Pt1 Pd1 /C and Ru@Pt1 Pd2 /C catalyst at 1000 s was 0.12, 0.10 and 0.08 mA cm2 , respectively. Ru@Pt2 Pd1 /C registered the highest current density, indicating enhanced electro-catalytic activity relative to Pd/C and the optimized Pt:Pd ratio amongst Ru@Ptx Pdy /C catalysts.

Current density / mA cm

-2

H. Gao et al. / Electrochimica Acta 56 (2011) 20242030

2029

80

0.5

Current density / mA cm

60

Ru@Pt1Pd1/C Ru@Pt2Pd1/C

Current density / mA cm

Pt2Pd1/C

-2

Pt/C Ru@Pt1Pd2/C

-2

A
0.4

40

Pd/C

0.3

0.2

20

Ru@Pt2Pd1/C Ru@Pt1Pd1/C Ru@Pt1Pd2/C

0.1

Pd/C
0.0 0.2 0.4 0.6 0.8 1.0 1.2
0.0 0 200 400 600 800 1000

Potential / V vs. NHE


Pt/C Pt2Pd1/C Ru@Pt2Pd1/C Ru@Pt1Pd1/C Ru@Pt1Pd2/C

t/s

Current density / mA cm

-2

600

Fig. 6. Chronoamperograms of the catalysts at 0.3 V in 2.0 M HCOOH + 0.50 M H2 SO4 at room temperature (electrode rotation speed: 600 rpm).

450

0.8
Ru@Pt2Pd1/C

10
Pt2Pd1/C

0.7

Power density / mW cm

300

Pt/C

0.6

Voltage / V

0.5 0.4 0.3 0.2

150

0 0.0 0.2 0.4 0.6 0.8 1.0 1.2

Potential / V vs. NHE


Pt/C Pt2Pd1/C Ru@Pt1Pd2/C Ru@Pt1Pd1/C Ru@Pt2Pd1/C

2 0.1

-2

400
Current density / mA cm
-2

0.0 0 10 20
-2

0 30

Current density / mA cm

300

Fig. 7. The polarization curves and power density curves of single cells with Ru@Pt2 Pd1 /C (circle), Pt2 Pd1 /C (square) and JM Pt/C (triangle). Fuel: 5 M formic acid; oxidant: air under atmospheric pressure.

200

with Pt2 Pd1 /C and 3.5 times higher than that of the MEA prepared with JM Pt/C. The higher single-cell performance of Ru@Pt2 Pd1 /C conrmed the above-discussed half cell evaluation results. 4. Conclusions

100

0 0.0 0.2 0.4 0.6 0.8 Potential / V vs. NHE 1.0 1.2

Fig. 5. Linear sweep voltammograms of Ru@Ptx Pdy /C, Pt/C, Pd/C and Pt2 Pd1 /C catalysts from 0.0 V to 1.2 V (A), from 1.2 to 0.0 V (B) and cyclic voltammograms of Ru@Ptx Pdy /C, Pt/C, Pd/C and Pt2 Pd1 /C catalysts (C) (electrolyte: N2 -purged 0.50 M H2 SO4 + 2.0 M HCOOH; sweep rate: 30 mV s1 , at room temperature).

Fig. 7 shows the single cell performance prepared with Ru@Pt2 Pd1 /C as the anode catalyst. For a fair comparison, the polarization curves and power density curves of single cell using Pt2 Pd1 /C and JM Pt/C as the anode catalysts are also shown. It can be seen that the cell using Ru@Pt2 Pd1 /C as anode catalyst gives a peak power density of 6.2 mW cm2 , which is higher than that with Pt2 Pd1 /C (5.5 mW cm2 ) or JM Pt/C (5.0 mW cm2 ) in the anode. The current density of the MEA with Ru@Pt2 Pd1 /C as the anode catalyst is 7.5 mA cm2 , over 30% higher than that of the MEA prepared

Carbon supported coreshell structured Ru@Ptx Pdy /C catalysts have been developed by deposition of Ptx Pdy on the pre-formed Ru/C catalysts. TEM observation found that the particle population of Ru@Ptx Pdy /C was similar to that of the pre-formed Ru/C except for particle grown-up, an indication of coreshell structure formation. The catalytic activity toward formic acid oxidation of the catalysts was evaluated by half-cell evaluation (cyclic voltammetry and chronoamperometric) and single cell measurements. It was found that Ru@Ptx Pdy /C displayed enhanced catalytic activity, improved stability and enhanced cell performance. Acknowledgments We would like to thank the National Scientic Foundation of China (NSFC Project Nos. 20673040 and 20876062), the Ministry of Science and Technology of China (No. 2009AA05Z119) and the Fundamental Research Funds for the Central Universities of South China University of Technology for nancial support of this work.

2030

H. Gao et al. / Electrochimica Acta 56 (2011) 20242030 [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] F.S. Thomas, R.I. Masel, Surf. Sci. 573 (2004) 169. W. Zhou, J.Y. Lee, Electrochem. Commun. 9 (2007) 1725. S. Wang, N. Kristian, S. Jiang, X. Wang, Electrochem. Commun. 10 (2008) 961. Y.N. Wu, S.J. Liao, Y.L. Su, J.H. Zeng, D. Dang, J. Power Sources 195 (2010) 6459. J. Jiang, A. Kucernak, J. Electroanal. Chem. 630 (2009) 10. L.M. Xu, S.J. Liao, L.J. Yang, Z.X. Liang, Fuel Cells 9 (2009) 101. J.H. Zeng, J.Y. Lee, W. Zhou, J. Power Sources 159 (2006) 509. R. Wang, H. Li, H. Feng, H. Wang, Z. Lei, J. Power Sources 195 (2010) 1099. S. Alayoglu, A.U. Nilekar, M. Mavrikakis, B. Eichhorn, Nat. Mater. 7 (2008) 333. W. Wang, R. Wang, S. Ji, H. Feng, H. Wang, Z. Lei, J. Power Sources 195 (2010) 3498. R. Chetty, S. Kundu, W. Xia, M. Bron, W. Schuhmann, V. Chirila, W. Brandl, T. Reinecke, M. Muhler, Electrochim. Acta 54 (2009) 4208. H. Gao, S. Liao, J. Zeng, Y. Xie, J. Power Sources 196 (2011) 54. H. Yang, N. Alonso-Vante, C. Lamy, D.L. Akins, J. Electrochem. Soc. 152 (2005) A704. X. Li, J. Liu, W. He, Q. Huang, H. Yang, J. Colloid Interface Sci. 344 (2010) 132. W. Li, P. Haldar, Electrochem. Commun. 11 (2009) 1195. S. Wang, N. Kristian, S. Jiang, X. Wang, Electrochem. Commun. 10 (2008) 961. Z. Zhang, J. Ge, L. Ma, J. Liao, T. Lu, W. Xing, Fuel Cells 9 (2009) 114.

References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] C. Rice, S. Ha, R.I. Masel, A. Wieckowski, J. Power Sources 115 (2003) 229. X. Wang, J.M. Hu, I.M. Hsing, J. Electroanal. Chem. 562 (2004) 73. J. Willsau, J. Heitbaum, Electrochim. Acta 31 (1986) 943. W. Chen, J. Kim, S. Sun, S. Chen, Phys. Chem. Chem. Phys. 8 (2006) 2779. M.D. Maci, E. Herrero, J.M. Feliu, J. Electroanal. Chem. 554 (2003) 25. S. Park, Y. Xie, M.J. Weaver, Langmuir 18 (2002) 5792. J.B. Xu, T.S. Zhao, Z.X. Liang, J. Phys. Chem. C 112 (2008) 17362. L.J. Zhang, Z.Y. Wang, D.G. Xia, J. Alloys Compd. 426 (2006) 268. S. Ha, R. Larsen, Y. Zhu, R.I. Masel, Fuel Cells 4 (2004) 337. S. Ha, R. Larsen, Y. Zhu, R.I. Masel, J. Power Sources 144 (2005) 28. L. Zhang, Y. Tang, J. Bao, T. Lu, C. Li, J. Power Sources 162 (2006) 177. K. Persson, A. Ersson, K. Jansson, N. Iverlund, S. Jrs, J. Catal. 231 (2005) 139. P. Waszczuk, T.M. Barnard, C. Rice, R.I. Masel, A. Wieckowski, Electrochem. Commun. 4 (2002) 599. [14] M. Arenz, V. Stamenkovic, T.J. Schmidt, K. Wandelt, P.N. Ross, N.M. Markovic, Phys. Chem. Chem. Phys. 5 (2003) 4242. [15] P.K. Babu, H.S. Kim, J.H. Chung, E. Oldeld, A. Wieckowski, J. Phys. Chem. B 108 (2004) 20228.

S-ar putea să vă placă și